Next Article in Journal
Non-Bosonic Damping of Spin Waves in van der Waals Ferromagnetic Monolayers
Next Article in Special Issue
Rational Construction of Nano-Scaled FeOOH/NiFe-LDH for Efficient Water Splitting
Previous Article in Journal
Research Status of Agricultural Nanotechnology and Its Application in Horticultural Crops
Previous Article in Special Issue
Application of Nanocomposites in Covalent Organic Framework-Based Electrocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

S-Doped FeOOH Layers as Efficient Hole Transport Channels for the Enhanced Photoelectrochemical Performance of Fe2O3

by
Yanhong Zhou
,
Yiran Zhang
,
Boyang Jing
,
Xiaoyuan Liu
and
Debao Wang
*
Key Laboratory of Inorganic Synthetic and Applied Chemistry, College of Chemistry and Molecular Engineering, Qingdao University of Science and Technology, Qingdao 266042, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(10), 767; https://doi.org/10.3390/nano15100767
Submission received: 1 April 2025 / Revised: 16 May 2025 / Accepted: 19 May 2025 / Published: 20 May 2025

Abstract

:
Hematite (Fe2O3) has been accepted as a promising and potential photo(electro)catalyst. However, its poor carrier separation and transfer efficiency has limited its application for photoelectrocatalytic (PEC) water oxidation. Herein, a S-doped FeOOH (S:FeOOH) layer was rationally designed and grown on Fe2O3 to construct a S:FeOOH/Fe2O3 composite photoanode. The obtained S:FeOOH/Fe2O3 photoanodes were fully characterized. The surface injection efficiency for Fe2O3 was then significantly increased with a high ηsurface value of 92.8%, which increases to 2.98 times for Fe2O3 and 2.16 times for FeOOH/Fe2O3, respectively. With 2.43 mA cm‒2 at 1.23 V, the optimized S:FeOOH/Fe2O3 photoanode was entrusted with a higher photocurrent density. The onset potential for S:FeOOH/Fe2O3 cathodically shifts 70 mV over Fe2O3. The improved PEC performance suggests that the S:FeOOH layer acts as ultrafast transport channels for holes at the photoanode/electrolyte interface, suppressing surface charge recombination. A Z-scheme band alignment between Fe2O3 and S:FeOOH was deduced from the UV–Vis and UPS spectra to promote charge transfer. This method provides an alternative for the construction of photoanodes with enhanced PEC water splitting performance.

Graphical Abstract

1. Introduction

Photoelectrochemical (PEC) water splitting using semiconductor photocatalytic electrodes has become an appealing technology for converting solar light into renewable hydrogen production to address energy-related and environmental problems [1,2]. The bottleneck of this technology is to obtain semiconductor photoelectrodes with a high photo-response and high catalytic activity to achieve efficient hydrogen production by water splitting [3,4]. To date, various semiconductor photoelectrode materials have been extensively investigated, including TiO2 [5,6], BiVO4 [7,8], α-Fe2O3 [9], WO3 [10], CdS [11], and so on [12]. Among numerous photocatalysts, hematite (Fe2O3) stands out as an excellent candidate for PEC water decomposition because of its moderate band-gap energy (1.9–2.2 eV), low cost, and excellent chemical stability [13,14,15,16]. However, the limited hole transmission distance (2−4 nm) and poor electron-hole carrier mobility in Fe2O3 hinder its efficient charge transfer and transport, mainly due to the significant hole recombination [17]. To address these crucial challenges, numerous studies have suggested various effective approaches, including morphology control [18,19], heterojunction construction [20,21,22,23], the incorporation of metal dopants [24,25,26,27], and cocatalyst deposition [28,29]. These approaches can effectively reduce photoanode overpotential while inducing an internal electric field, thereby improving band alignment [30,31,32,33,34].
Alternatively, FeOOH has attracted considerable attention as a promising oxygen evolution catalyst due to its environmental friendliness, non-toxicity, and abundance [35,36,37]. It has been examined as a hole transfer layer or as a passivation layer to mitigate the recombination of electron-hole pairs [38,39]. However, the low electrical conductivity of pure FeOOH remains a significant limitation that negatively affects its catalytic performance [40]. To improve efficiency, further efforts have been devoted to strategies such as metal/non-metal doping [41,42], the introduction of oxygen vacancies [43,44], and phase structure transitions [45,46]. In particular, elemental doping has been widely accepted to improve catalytic activity by modulating the electron structures of the surrounding metallic active species [47,48]. For example, Kim et al. discovered that the doping of Mn in iron oxyhydroxide (FeOOH) results in an increase in intrinsic activity by changing the electron structures of FeOOH [49]. Wang et al. stated that the surface injection efficiency of BiVO4 photoanodes can be significantly enhanced by growing a Ni-doped FeOOH layer on the anode [50]. Recently, He et al. reported that the doping of FeOOH with S effectively improves the charge transfer capacity of FeOOH and the surface catalytic reaction kinetics in the synthesized S-FeOOH/BiVO4 photoanode [51]. However, the current explanations are still unclear, and the enhancement in photocurrent signals by element doping and co-catalyst modifying requires more in-depth investigations. For example, the charge transfer mechanism between S-FeOOH and Fe2O3 remains unclear.
Herein, we report the preparation of S:FeOOH/Fe2O3 photoanodes with a sulfur-doped FeOOH layer modified on Fe2O3. The optimized S:FeOOH/Fe2O3 photoanode achieved a photocurrent density of 2.43 mA cm−2 at 1.23 V, showing a significant increase in surface injection efficiency to 92.8% and good stability. This performance enhancement is attributed to the influence of sulfur on the electronic structure of FeOOH, which acts as an efficient channel for hole transport at the electrode/electrolyte interface. In addition, the S:FeOOH layer reduces the onset potential by 70 mV compared to Fe2O3 when acting as oxygen evolution catalysts. Furthermore, from the results of the ultraviolet photoelectron spectra (UPS) and UV–Vis spectra, a Z-scheme band alignment between S:FeOOH and Fe2O3 has been deduced, allowing for improved charge carrier transfer.

2. Materials and Methods

2.1. Synthesis of the Photoanodes

The synthesis procedure for the S:FeOOH/Fe2O3 photoanodes is illustrated in Scheme 1. Firstly, a piece of fluorine-doped tin oxide (FTO) glass slide substrate (2.5 cm × 1 cm) was pre-treated by washing with acetone, ethanol, and deionized water. For the synthesis of the Fe2O3 photoanode, 0.15 mmol Fe(NO3)3·9H2O and 1.0 mmol oxalic acid precursor were added to 60 mL of deionized water and dissolved to form a transparent light yellow solution. It was then transferred to an 80 mL Teflon-lined autoclave and heated at 180 °C for 8 h in an oven. After natural cooling to ambient temperature, the resulting photoanode was taken out, rinsed repeatedly with deionized water and absolute ethanol, dried naturally, and finally annealed in a muffle furnace at 550 °C in air for 2 h.
For the synthesis of the FeOOH/Fe2O3 photoanode, 0.02 g Fe(NO3)3·9H2O and 0.01 g urea were first dissolved in 30 mL of deionized water in a 50 mL Teflon-lined autoclave. After immersing the pre-prepared Fe2O3 photoanode in the solution, the autoclave was sealed and heated at 90 °C for 90 min. After the reaction, the photoanode was rinsed three times with deionized water and absolute ethanol and then dried at 60 °C.
For the synthesis of the S:FeOOH/Fe2O3 photoanode, the pre-prepared FeOOH/Fe2O3 photoanode was immersed in 15 mL ethanol solution containing 0.3 g thioacetamide and kept at 95 °C for 3 h. After the reaction, the resulting electrode was processed in the same way as the FeOOH/Fe2O3 photoanode.

2.2. PEC Measurements

A CHI760E electrochemical workstation was used for all PEC measurements in a three-electrode system. The as-prepared photoanode (1 × 1 cm2 in size) was used as the working electrode. The reference electrode was a saturated Ag/AgCl electrode (saturated with KCl). The counter electrode was a Pt wire. Photocurrent density–potential (J-V), photocurrent density–time (J-t), and linear sweep voltammetry (LSV) curves were recorded in 1 M KOH at 1.23 V under a 300 W Xe-lamp (100 mW·cm−2). The cyclic voltammograms (CV) were recorded in a non-Faradaic region (0.90–1.00 V). In this research, the measured potentials were converted to RHE: V (vs. RHE) = V (vs. Ag/AgCl) + 0.197 + 0.059 pH. The EIS (electrochemical impedance spectroscopy) experiments were performed at 0.2 V vs. RHE in a frequency range of 0.1 Hz–100 kHz.
The incident photon-to-current conversion efficiency (IPCE) and the applied bias photon-to-current efficiency (ABPE) were conducted on a Modulab XM PhotoEchem workstation (Solartron Analytical, Farnborough, UK) in the 350–700 nm range with a 1.23 V bias, illuminated using a 300 W Xe lamp. The IPCE values were calculated using the following equation: IPCE (%) = 1240 × Jph/(P × λ) × 100%. In this equation, λ is the wavelength, Jph(λ) is the photocurrent density measured at a given wavelength, and P(λ) is the intensity of the incident light. The ABPE values were calculated using the following formula [52]:
ABPE   ( % ) = J × ( 1.23 V R H E ) P l i g h t × 100 %
where VRHE is the applied bias voltage, J denotes the photocurrent density at the applied bias voltage, and Plight represents the power intensity taking a value of 100 mW·cm−2.
The details of the material characterization are presented as the Supporting Information.

3. Results and Discussions

3.1. Material Characterization

Following the synthesis process illustrated in Scheme 1, S:FeOOH/Fe2O3 photoanodes were obtained. Figure 1a shows the SEM image of the Fe2O3 photoanode. It is obvious that a rough and porous surface rather than a dense and smooth surface was formed with irregular Fe2O3 nanoparticles. The XRD pattern of the Fe2O3 photoanode in Figure 1d indicates the presence of the hexagonal-phase hematite α-Fe2O3 according to the standard diffraction pattern in PDF#33-0664 together with the tetragonal phase SnO2 of the FTO (PDF#46-1088), confirming the formation of the Fe2O3/FTO composite. Subsequently, an FeOOH layer was deposited on the Fe2O3 photoanode using the hydrothermal deposition method. The SEM image presented in Figure 1b reveals the surface morphology of the FeOOH/Fe2O3 anode. After the FeOOH layer is loaded, the surfaces of the Fe2O3 nanoparticles becomes much rougher in comparison with that of the Fe2O3 photoanode. This indicates that the Fe2O3 photoanode is completely covered with the FeOOH layer through the homogenous nucleation and growth of FeOOH nanoparticles during the hydrothermal process [40]. Subsequently, S-loading was carried out via a solvothermal approach. The SEM image of the S:FeOOH/Fe2O3 photoanode is shown in Figure 1c, which indicates no obvious changes in the surface morphology in comparison to that of FeOOH/Fe2O3. However, the S-doping results in a much rougher surface for the S:FeOOH/Fe2O3 photoanode. The XRD patterns of FeOOH/Fe2O3 and S:FeOOH/Fe2O3 photoanodes are shown in Figure 1d. The diffraction peaks correspond to those of the α-Fe2O3, but no obvious peaks correspond to FeOOH, indicating the very low loading amount and/or amorphous nature of S:FeOOH.
The TEM image in Figure 2a shows a significant dark and light contrast, further indicating that the photoanode material was constructed with irregular S:FeOOH/Fe2O3 nanoparticles. The HRTEM image in Figure 2b reveals the microstructure of the S:FeOOH/Fe2O3 nanoparticles composed of crystalline Fe2O3 nanoparticles and amorphous S:FeOOH nanoparticles, indicating the formation of the FeOOH layer at the surface of the Fe2O3 anode. The lattice spacing of 0.25 nm corresponds to the (110) planes of α-Fe2O3 according to the data of PDF#33-0664, which is in agreement with the XRD results. To see the S element locations at the atomic level, the STEM HADDF image and the corresponding EDS element mappings of Fe, O, and S are presented in Figure S1. It is evident that S element homogenously distributes in the whole detection region, indicating that sulfur has been successfully doped into the S:FeOOH/Fe2O3 anode.
FT-IR and Raman spectra were recorded to further identify the composition of the electrodes. As shown in Figure S2a, the FT-IR spectra reveal the characteristic stretching vibrations of –OH groups and Fe–O bonds, as well as the Fe–OH bending vibrations. Compared to FeOOH/Fe2O3, the vibrations around 1110 cm−1 become significantly more serrated, corresponding to the S–O vibration [53], and the absorption bands of S:FeOOH/Fe2O3 exhibit a slight red shift. These results indicate that S2− was successfully doped into FeOOH/Fe2O3 [54]. The Raman spectra of the FeOOH/Fe2O3/FTO and S:FeOOH/Fe2O3/FTO anodes are presented in Figure S2b. Both spectra show the characteristic modes of FeOOH and Fe2O3 [55]. After S doping, the Raman peaks of the S:FeOOH/Fe2O3/FTO anode become broader and slightly weaker, implying decreased particle sizes of the S-doped FeOOH/Fe2O3 nanoparticles. This further indicates that S doping altered the microstructure of the FeOOH/Fe2O3 nanoparticles [54].
The chemical state of the relevant surface elements was characterized by means of XPS. As illustrated in Figure 3a, the high-resolution XPS spectra of Fe 2p exhibit two prominent peaks at approximately 710.8 eV and 724.4 eV, associating with Fe 2p3/2 and Fe 2p1/2, respectively. In addition, shakeup satellite peaks can be observed at around 718.2 eV and 732.9 eV. In comparison with the pristine Fe2O3, the S:FeOOH/Fe2O3 anode exhibits a positive shift of 0.5 eV in the binding energy of Fe 2p. The binding energy shift of Fe 2p after S doping indicates that sulfur should be successfully doped into FeOOH and that the electronic modulation occurs in S:FeOOH/Fe2O3 [51,56]. The O 1s spectra in Figure 3b reveals an increased peak at 532.4 eV, corresponding to surface-absorbed hydroxyl (Fe-OH) groups and a lattice oxygen peak (OL) around 529.8 eV. It is evident that the relative intensity of the -OH peak for S:FeOOH/Fe2O3 increases significantly. The peak area ratio of the fitted -OH peak to OL peak increases from 0.69:1 for S:FeOOH/Fe2O3 to 6.1:1 for S:FeOOH/Fe2O3, thereby suggesting the successful deposition of the FeOOH layer onto the Fe2O3 film [57]. Furthermore, the OL peak for the S:FeOOH/Fe2O3 sample shifts to a higher energy level, suggesting the formation of an internal electric field at its interface with Fe2O3, thereby promoting efficient charge transport across the interface of S:FeOOH/Fe2O3 [57]. Figure 3c shows the S 2p spectrum. The fitted peaks around 165 eV are assigned to S 2p3/2 and S 2p1/2 of metal sulfur bonding energy, while those around 169 eV correspond to oxidized sulfur [51]. For comparison, the S 2p spectrum for the Fe2O3 anode was also recorded and is presented in Figure 3c. It is evident that no obvious S 2p XPS signals have been recorded, which provides evidence for sulfur doping into the FeOOH layer via the solvothermal method. The presence of these S peaks provides further confirmation of the efficacy of sulfur doping into the FeOOH layer via the solvothermal method.

3.2. PEC Results

The PEC properties of Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 photoanodes were evaluated using 1 M KOH. The optimized reaction time for S-doping was determined to be 3 h according to the LSV curves presented in Figure 4a. The LSV curves in Figure 4b show that no obvious photocurrent was recorded at 1.23 V in the dark for any of the anodes. Under irradiation, pure Fe2O3 achieves only 0.37 mA·cm−2 photocurrent density at 1.23 V, which increases to 0.87 mA·cm−2 after the FeOOH layer is deposited on the Fe2O3 surface. The S:FeOOH/Fe2O3 anode achieves an even higher photocurrent density of 2.43 mA·cm−2. This represents a 5.6-fold increase over pure Fe2O3 and a 1.8-fold increase over FeOOH/Fe2O3. The Tafel plots in Figure 4c for Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 were found to be 156.9 mV, 83.8 mV, and 53.0 mV dec−1, respectively. These results directly imply that the S:FeOOH layer makes a huge contribution to the water oxidation kinetics of the S:FeOOH/Fe2O3 photoelectrodes. Figure 4d displays the Butler plots (J2 vs. V plots) of different photoanodes. It is evident that the onset potential of Fe2O3 exhibits a cathodic shift of approximately 70 mV, decreasing from an initial value of 0.93 V to 0.86 V, indicating that the loading of the S:FeOOH layer can effectively improve the PEC activity of the Fe2O3 anode [57].
Figure 5a displays the photocurrent responses of different photoanodes. It indicates that the steady-state photocurrents during each light-on period exhibit excellent agreement with the LSV data. The ABPE plot (Figure 5b) shows that the maximum ABPE for S:FeOOH/Fe2O3 is 0.24%, which is eight times higher than the 0.03% recorded for Fe2O3. The IPCE for Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 at 1.23 V vs. RHE is presented in Figure 5c. It shows that S:FeOOH/Fe2O3 achieves the highest IPCE value of 31.5%, exceeding the values of bare Fe2O3 (12.3%) and FeOOH/Fe2O3 (17.7%). Band gaps can be obtained from IPCE using a Tauc plot analysis of (IPCE % × hν)1/2 versus hν (Figure S3) in electrochemical noise mode [33]. Figure S3 shows that Fe2O3 has a band gap of 2.1 eV and the value for S:FeOOH is 1.86 eV. In addition, as shown in Figure 5d, S:FeOOH/Fe2O3 exhibits excellent stability for a period of 2 h with no decrease in photocurrent density, indicating an excellent PEC water oxidation stability.

3.3. Charge Transfer Kinetics

To investigate the surface and bulk charge recombination in different photoanodes, the surface charge separation efficiency (ηsurface) and the bulk charge separation efficiency (ηbulk) were evaluated with 0.5 M Na2SO3 as a hole scavenger. Figure 6a shows the LSV curves of different photoanodes. When Na2SO3 is added to the electrolyte solution, the photogenerated holes can immediately participate in the oxidation of the water molecules arriving at the photoanode/electrolyte interface. As a result, the surface hole recombination is effectively suppressed. Due to the rapid oxidation kinetics of SO32− species, it can be assumed that ηsurface is 100% for SO32− oxidation [34]. The ηsurface and ηbulk are calculated using the equations ηsurface = J/JNa2SO3 and ηbulk = JNa2SO3/Jabs, where JNa2SO3 and J represent, respectively, the photocurrents with and without a Na2SO3 scavenger, respectively, and Jabs denotes the theoretical photocurrent density assuming that the absorbed photons completely convert into current, which is referred to as 11.4 mA cm−2 [58,59]. As presented in Figure 6b, the ηbulk values for Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 are 10.4%, 18.2%, and 23.0%, respectively, at 1.23 V. It is evident that a significant improvement has been achieved in the bulk charge separation efficiency with the introduction of S:FeOOH. In Figure 6c, the ηsurface values of Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 at 1.23 V are 31.1%, 42.9%, and 92.8%, respectively. The ηsurface value of S:FeOOH/Fe2O3 is 2.98 times and 2.66 times higher than those of the Fe2O3 and FeOOH/Fe2O3 anodes, respectively. The significant improvement can be attributed to the rapid injection of surface charges into the oxide water molecules via the S:FeOOH layer. Consequently, the S:FeOOH layer can contribute a more effective catalytic effect to the surface reactions than FeOOH alone.
The cyclic voltammetry (CV) curves of different photoanodes were recorded, and the results are presented in Figure 7a–c, from which the corresponding electrochemical double-layer capacitances (Cdl) can be deduced. As shown in Figure 7d, the Cdl value for S:FeOOH/Fe2O3 (360 F·cm−2) is significantly higher than that of FeOOH/Fe2O3 (134 F·cm−2) and Fe2O3 (46 F·cm−2). Then, the electrochemical active surface area (ECSA) of the photoanodes can be evaluated according to the proportional relationship of ECSA and Cdl. Therefore, the S:FeOOH/Fe2O3 photoanode should have a significantly higher ECSA value, thereby increasing its intrinsic electrochemical active surface area at the catalyst/liquid interface.
The dynamics of the charge transfer at the electrolyte/photoelectrode interface was also examined using EIS techniques. The recorded Nyquist curves are presented in Figure 8a. It is evident that all photoanodes exhibit a characteristic arc with varying radii over the frequency range. The smaller the impedance radius, the better the conductivity and the lower the interfacial charge transfer resistance. Apparently, the S:FeOOH/Fe2O3 photoanode exhibits a minimum value of arc radii, indicating improved charge transfer performance. The observed reduction in impedance can be attributed to an increase in the internal electric field strength within Fe2O3, resulting from the introduction of the S:FeOOH layer. This enhancement promotes both electron and hole transport while simultaneously reducing surface recombination rates. Specifically, the S:FeOOH layer acts as an efficient hole transport channel, efficiently extracting photogenerated holes from Fe2O3 and accelerating the charge transfer.
Mott–Schottky (M-S) curves were carried out to further investigate the charge transfer kinetics. As illustrated in Figure 8b, the flat band potential (Efb) for pristine Fe2O3 is estimated to be approximately 0.59 V vs. RHE, and it exhibits a slight cathodic shift to approximately to 0.50 V and 0.41 V for FeOOH/Fe2O3 and S:FeOOH/Fe2O3, respectively. Additionally, the charge carrier density (Nd) was calculated from the M-S curve slope using the equation [60] Nd = (2/e0ɛɛ0)[d(1/C2)/dV]−1, where Nd represents the charge carrier density, e0 is the electron charge, ɛ is the dielectric constant of Fe2O3, ɛ0 is the vacuum permittivity, and V is the applied electrode potential. Therefore, the charge carrier densities were calculated to be 1.31 × 1021 cm−3, 1.75 × 1021 cm−3, and 4.12 × 1021 cm−3 for Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3, respectively. The carrier density of S:FeOOH/Fe2O3 is 2.4 times that of FeOOH/Fe2O3, proving that the introduction of the S:FeOOH layer can efficiently accelerate the formation of charge carriers, thus increasing the conductivity of the electrodes and benefiting the fast hole transport and photochemical water oxidation.
The transient photovoltage (TPV) plots for Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3 photoelectrodes are presented in Figure 8c. It is clear that both Fe2O3 and FeOOH/Fe2O3 exhibit a pronounced negative TPV signal at 1.0 × 10−6 s, which can be attributed to the significant charge carrier recombination [61]. In contrast, the S:FeOOH/Fe2O3 photoanode shows a significant positive TPV signal, indicating the superior separation efficiency of the photogenerated charges. This observation is consistent with the EIS results shown in Figure 8a, confirming the enhanced charge separation capabilities and reduced hole recombination of the S:FeOOH/Fe2O3 structure. Consequently, the S:FeOOH layer functions as an efficient hole transport channel between itself and the light-absorbing layer, facilitating efficient charge carrier separation.

3.4. Mechanism Investigation

To investigated the PEC mechanism of S:FeOOH/Fe2O3 for enhanced PEC performance, and we analyzed the valence band structures of Fe2O3 and S:FeOOH photoanodes using UPS and UV–Vis spectra. Figures S4 and S5 present the UV–Vis absorption spectra of Fe2O3, FeOOH/Fe2O3, and S:FeOOH/Fe2O3, which show no significant changes in light absorption. This suggests that the improved photoelectrochemical performance of S:FeOOH/Fe2O3 is not due to the changes in light absorption, but rather due to other factors such as improved charge transfer and catalytic activity [51]. The energy gaps (Eg) are derived from the Tauc plots (Figure S5) from the UV–Vis absorption spectra. The results are in agreement with the Tauc plot analysis from IPCE. The secondary electron cut-off energies in Figure 9a,c are used to determine the work function, while the binding energy region shown in Figure 9b,d is used to determine the valence band maxima (VBM) for both the Fe2O3 and S:FeOOH anodes. Table S1 summarizes the values of the work functions, VBM, conduction band minima (CBM), and Eg values for Fe2O3 and S:FeOOH.
As shown in Table S1, the calculated band alignments for Fe2O3 and S:FeOOH exhibit a Z-scheme configuration, which is illustrated in Figure 10a, while Figure 10b schematically illustrates the charge transfer mechanism within the S:FeOOH/Fe2O3 photoanode. With illumination, electron-hole pairs are excited in both materials. Specifically, photogenerated electrons transfer from the CB of S:FeOOH to the VB of Fe2O3. Simultaneously, photogenerated holes migrate from the VBM of S:FeOOH into water to produce O2, and photogenerated electrons migrate from the CBM of Fe2O3 to be transported into the counter electrode to produce H2. This process effectively suppresses the strong surface recombination on Fe2O3 by establishing an efficient hole transport channel through the overlying S:FeOOH layer.

4. Conclusions

In conclusion, an S-doped FeOOH layer has been rationally designed and deposited on the surface of the Fe2O3 anodes using hydrothermal and subsequent solvothermal methods. The catalyst consists of crystalline nanoparticles and an amorphous S-doped FeOOH layer. The Z-scheme band alignment between Fe2O3 and S:FeOOH effectively increases the electron-hole pair separation efficiency. The S:FeOOH layer not only significantly increases the ECSA of the Fe2O3 anode, but also serves as an efficient hole transport channel at the electrolyte/photoanode interface, thereby enhancing the charge carrier transfer efficiency. Accordingly, the S:FeOOH/Fe2O3 photoanode achieved a significantly higher photocurrent density of 2.43 mA cm−2 at 1.23 V with a good stability, which is 5.6 times higher than pristine Fe2O3 and 1.8 times higher than FeOOH/Fe2O3. Its onset potential also has a 70 mV cathodic shift compared to Fe2O3. Moreover, the S:FeOOH/Fe2O3 photoanode achieves a surface efficiency (ηsurface) of 92.8%, which is 2.98 times higher than Fe2O3 and 2.66 times higher than FeOOH/Fe2O3. This simple yet effective approach offers a promising strategy for the deposition of oxygen evolution catalysts with improved performance in PEC water splitting H2 production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15100767/s1, Material characterizations; Figure S1: HADDF image and element mappings; Figure S2: FT-IR and Raman spectra; Figure S3: Band gaps of different anodes from IPCE; Figures S4 and S5: UV-Vis absorption spectra and Tauc plots of different photoanodes; Table S1: The values of work function, EVBM, ECBM and Eg of Fe2O3 and S:FeOOH photoanodes.

Author Contributions

Conceptualization, Y.Z. (Yanhong Zhou); methodology, Y.Z. (Yanhong Zhou); validation, Y.Z. (Yiran Zhang) and B.J.; investigation, Y.Z. (Yanhong Zhou), B.J. and X.L.; writing—original draft preparation, Y.Z. (Yanhong Zhou); writing—review and editing, D.W.; supervision, D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 51872152) and the innovation and entrepreneurship training program for college student of Shandong province (S202410426032).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Pan, L.F.; Dai, L.L.; Burton, O.J.; Chen, L.; Andrei, V.; Zhang, Y.C.; Ren, D.; Cheng, J.; Wu, L.; Frohna, K.; et al. High carrier mobility along the [111] orientation in Cu2O photoelectrodes. Nature 2024, 628, 765–770. [Google Scholar] [CrossRef] [PubMed]
  2. Li, X.P.; Zhang, C.X.; Geng, J.F.; Zong, S.H.; Wang, P.Q. Photo(electro)catalytic Water Splitting for Hydrogen Production: Mechanism, Design, Optimization, and Economy. Molecules 2025, 30, 630. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, J.; Liu, K.; Liao, W.R.; Kang, Y.C.; Xiao, H.R.; Chen, Y.K.; Wang, Q.Y.; Luo, T.; Chen, J.; Li, H.; et al. Metal vacancies in semiconductor oxides enhance hole mobility for efficient photoelectrochemical water splitting. Nat. Catal. 2025, 8, 229–238. [Google Scholar] [CrossRef]
  4. Liu, S.Q.; Wu, L.; Tang, D.J.; Xue, J.; Dang, K.; He, H.B.; Bai, S.M.; Ji, H.; Chen, C.; Zhang, Y.; et al. Transition from sequential to concerted proton-coupled electron transfer of water oxidation on semiconductor photoanodes. J. Am. Chem. Soc. 2023, 145, 23849–23858. [Google Scholar] [CrossRef]
  5. Maurya, O.; Khaladkar, S.R.; Sinha, B.; Bhanage, B.M.; Deshmukh, R.R.; Kim, J.H.; Kalekar, A. Effective transformation of hydrothermally grown TiO2 nanorods to nanotube arrays for improved PEC hydrogen evolution. Electrochim. Acta 2023, 471, 143391. [Google Scholar] [CrossRef]
  6. Arunachalam, P.; Amer, M.S.; AlOraij, H.A.; Al-Mayouf, A.M.; Hezam, M.; Al-Shalwi, M. Boosting the Photoelectrochemical Water Oxidation Performance of TiO2 Nanotubes by Surface Modification Using Silver Phosphate. Catalysts 2022, 12, 1440. [Google Scholar] [CrossRef]
  7. Fu, H.; Qi, Q.; Li, Y.; Pan, J.; Zhong, C. Oxygen-Vacancy-Induced Enhancement of BiVO4 Bifunctional Photoelectrochemical Activity for Overall Water Splitting. Nanomaterials 2024, 14, 1270. [Google Scholar] [CrossRef]
  8. Zhao, H.; Wei, X.; Pei, Y.; Han, W. Enhancing Photoelectrocatalytic Efficiency of BiVO4 Photoanodes by Crystal Orientation Control. Nanomaterials 2024, 14, 1870. [Google Scholar] [CrossRef]
  9. Chen, D.; Liu, Z.F.; Zhang, S.C. Enhanced PEC performance of hematite photoanode coupled with bimetallic oxyhydroxide NiFeOOH through a simple electroless method. Appl. Catal. B 2020, 265, 118580. [Google Scholar] [CrossRef]
  10. Li, Y.; Mei, Q.; Liu, Z.J.; Hu, X.S.; Zhou, Z.H.; Huang, J.W.; Bai, B.; Liu, H.; Ding, F.; Wang, Q.Z. Fluorine-doped iron oxyhydroxide cocatalyst: Promotion on the WO3 photoanode conducted photoelectrochemical water splitting. Appl. Catal. B 2022, 304, 120995. [Google Scholar] [CrossRef]
  11. Li, W.L.; Geng, H.C.; Yao, L.; Cao, K.S.; Sheng, P.T.; Cai, Q.Y. Photoelectrocatalytic Hydrogen Generation Enabled by CdS Passivated ZnCuInSe Quantum Dot-Sensitized TiO2 Decorated with Ag Nanoparticles. Nanomaterials 2019, 9, 393. [Google Scholar] [CrossRef]
  12. Pech-Rodríguez, W.J.; Sahin, N.E.; Suarez-Velázquez, G.G.; Meléndez-González, P.C. Semiconductor-Based Photoelectrocatalysts in Water Splitting: From the Basics to Mechanistic Insights—A Brief Review. Materials 2025, 18, 1952. [Google Scholar] [CrossRef]
  13. Huang, Q.Y.; Zhao, Y.C.; Li, Y.D. Improved photoelectrochemical water splitting performance of Sn-doped hematite photoanode with an amorphous cobalt oxide layer. Int. J. Hydrog. Energ. 2024, 51, 1176–1183. [Google Scholar] [CrossRef]
  14. Xu, C.Y.; Wang, H.X.; Guo, H.Y.; Liang, K.; Zhang, Y.M.; Li, W.C.; Chen, J.; Lee, J.S.; Zhang, H. Parallel multi-stacked photoanodes of Sbdoped p–n homojunction hematite with near-theoretical solar conversion efficiency. Nat. Commun. 2024, 15, 9712. [Google Scholar] [CrossRef] [PubMed]
  15. Humayun, A.; Manivelan, N.; Prabakar, K. Charge Transfer in n-FeO and p-α-Fe2O3 Nanoparticles for Efficient Hydrogen and Oxygen Evolution Reaction. Nanomaterials 2024, 14, 1515. [Google Scholar] [CrossRef] [PubMed]
  16. Baldovi, H.G. Optimization of α-Fe2O3 Nanopillars Diameters for Photoelectrochemical Enhancement of α-Fe2O3-TiO2 Heterojunction. Nanomaterials 2021, 11, 2019. [Google Scholar] [CrossRef] [PubMed]
  17. Suman; Singh, S.; Ankita; Kumar, A.; Kataria, N.; Kumar, S.; Kumar, P. Photocatalytic activity of α-Fe2O3@CeO2 and CeO2@α-Fe2O3 core-shell nanoparticles for degradation of Rose Bengal dye. J. Environ. Chem. Eng. 2021, 9, 106266. [Google Scholar] [CrossRef]
  18. Sun, J.; Xia, W.; Zheng, Q.; Zen, X.; Liu, W.; Li, G.; Wang, P. Increased active sites on irregular morphological α-Fe2O3 nanorods for enhanced photoelectrochemical performance. ACS Omega 2020, 5, 12339–12345. [Google Scholar] [CrossRef]
  19. Yan, K.; Qiu, Y.; Xia, S.; Gon, J.; Zhao, S.; Xu, J. Self-driven hematite-based photoelectrochemical water splitting cells with three-dimensional nanobowl heterojunction and high-photovoltage perovskite solar cells. Mater. Today Energy 2017, 6, 128–135. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Huang, Y.; Zhu, S.S.; Liu, Y.Y.; Zhang, X.; Wagn, J.J.; Braun, A. Covalent S-O bonding enables enhanced photoelectrochemical performance of Cu2S/Fe2O3 heterojunction for water splitting. Small 2021, 17, 2100320. [Google Scholar] [CrossRef]
  21. Ma, J.; Wang, Q.; Li, L.; Zong, X.; Sun, H.; Tao, R.; Fan, X. Fe2O3 nanorods/CuO nanoparticles p-n heterojunction photoanode: Effective charge separation and enhanced photoelectrochemical properties. J. Colloid Interface Sci. 2021, 602, 32–42. [Google Scholar] [CrossRef] [PubMed]
  22. Masoumi, Z.; Tayebi, M.; Kolaei, M.; Tayyeb, A.; Ryu, H.; Jang, J.I.; Lee, B.K. Simultaneous enhancement of charge separation and hole transportation in a W: α-Fe2O3/MoS2 photoanode: A collaborative approach of MoS2 as a heterojunction and W as a metal dopant. ACS Appl. Mater. Interfaces 2021, 13, 39215–39229. [Google Scholar] [CrossRef]
  23. Li, Y.; Wu, Q.; Chen, Y.; Zhang, R.; Li, C.; Zhang, K.; Li, M.; Lin, Y.; Wang, D.; Zou, X.; et al. Interface engineering Z-scheme Ti-Fe2O3/In2O3 photoanode for highly efficient photoelectrochemical water splitting. Appl. Catal. B 2021, 290, 120058. [Google Scholar] [CrossRef]
  24. Yoon, K.Y.; Park, J.; Jung, M.; Ji, S.G.; Lee, H.; Seo, J.H.; Kwak, M.J.; Il Seok, S.; Lee, J.H.; Jang, J.H. NiFeOx decorated Ge-hematite/perovskite for an efficient water splitting system. Nat. Commun. 2021, 12, 4309. [Google Scholar] [CrossRef] [PubMed]
  25. Zhou, Z.; Long, R.; Prezhdo, O.V. Why silicon doping accelerates electron polaron diffusion in hematite. J. Am. Chem. Soc. 2019, 141, 20222–20233. [Google Scholar] [CrossRef]
  26. Zhang, S.; Zhang, S.; Leng, W.; Wu, D. Ti doped α-Fe2O3 electrodes for water oxidation. J. Alloys Compd. 2022, 927, 166975. [Google Scholar] [CrossRef]
  27. Nyarige, J.S.; Paradzah, A.T.; Krüger, T.P.J.; Diale, M. Mono-Doped and Co-Doped Nanostructured Hematite for Improved Photoelectrochemical Water Splitting. Nanomaterials 2022, 12, 366. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, J.; García-Rodríguez, R.; Cameron, P.; Eslava, S. Role of cobalt–iron (oxy) hydroxide (CoFeOx) as oxygen evolution catalyst on hematite photoanodes. Energy Environ. Sci. 2018, 11, 2972–2984. [Google Scholar] [CrossRef]
  29. Chang, Y.; Han, M.; Ding, Y.; Wei, H.; Zhang, D.; Luo, H.; Li, X.; Yan, X. Interface Engineering of CoFe-LDH Modified Ti:α-Fe2O3 Photoanode for Enhanced Photoelectrochemical Water Oxidation. Nanomaterials 2023, 13, 2579. [Google Scholar] [CrossRef]
  30. Wu, H.H.; Ding, H.H.; Ma, J.; Li, X.; Huai, S.S.; Zhang, C.M.; Li, P.; Xia, J.J.; Fang, X.L.; Wang, X.F. P-N homojunction and heteroatom active site engineering over Fe2O3 nanorods for highly efficient photoelectrochemical water splitting. Int. J. Hydrog. Energ. 2024, 88, 965–976. [Google Scholar] [CrossRef]
  31. Jeong, I.K.; Mahadik, M.A.; Hwang, J.B.; Chae, W.S.; Choi, S.H.; Jang, J.S. Lowering the onset potential of Zr-doped hematite nanocoral photoanodes by Al co-doping and surface modification with electrodeposited Co-Pi. J. Colloid Interface Sci. 2021, 581, 751–763. [Google Scholar] [CrossRef] [PubMed]
  32. Xia, W.W.; Zhang, R.; Chai, Z.C.; Pu, J.Y.; Kang, R.; Wu, G.Q.; Zeng, X.H. Synergies of Zn/P-co-doped α-Fe2O3 photoanode for improving photoelectrochemical water splitting performance. Int. J. Hydrog. Energ. 2024, 59, 22–29. [Google Scholar] [CrossRef]
  33. Ji, M.H.; Chen, Y.X.; Chen, R.; Li, K.X.; Zhao, H.-P.; Shi, H.Y.; Wang, H.L.; Jiang, X.; Lu, C.Z. A novel α-Fe2O3 photoanode with multilayered In2O3/Co–Mn nanostructure for efficient photoelectrochemical water splitting. Int. J. Hydrog. Energ. 2024, 51, 66–77. [Google Scholar] [CrossRef]
  34. Kim, N.; Ju, S.; Ha, J.; Choi, H.; Sung, H.; Lee, H. Hierarchical Co–Pi Clusters/Fe2O3 Nanorods/FTO Micropillars 3D Branched Photoanode for High-Performance Photoelectrochemical Water Splitting. Nanomaterials 2022, 12, 3664. [Google Scholar] [CrossRef] [PubMed]
  35. Ge, G.; Liu, M.; Liu, C.; Zhou, W.; Wang, D.; Liu, L.; Ye, J. Ultrathin FeOOH nanosheets as an efficient cocatalyst for photocatalytic water oxidation. J. Mater. Chem. A 2019, 7, 9222–9229. [Google Scholar] [CrossRef]
  36. Wang, T.; Gao, L.; Wang, P.; Long, X.; Chai, H.; Li, F.; Jin, J.J. Dual-doping in the bulk and the surface to ameliorate the hematite anode for photoelectrochemical water oxidation. J. Colloid Interface Sci. 2022, 624, 60−69. [Google Scholar] [CrossRef]
  37. Zhang, W.; Zhang, Y.; Miao, X.; Zhao, L.; Zhu, C. Deposition of FeOOH Layer on Ultrathin Hematite Nanoflakes to Promote Photoelectrochemical Water Splitting. Micromachines 2024, 15, 387. [Google Scholar] [CrossRef]
  38. Li, R.X.; Zhan, F.Q.; Wen, G.C.; Wang, B.; Qi, J.H.; Liu, Y.S.; Feng, C.C.; La, P.Q. Facile Synthesis of a Micro–Nano-Structured FeOOH/BiVO4/WO3 Photoanode with Enhanced Photoelectrochemical Performance. Catalysts 2024, 14, 828. [Google Scholar] [CrossRef]
  39. Lu, X.Y.; Ye, K.H.; Zhang, S.Q.; Zhang, J.; Yang, J.D.; Huang, Y.C.; Ji, H.B. Amorphous type FeOOH modified defective BiVO4 photoanodes for photoelectrochemical water oxidation. Chem. Eng. J. 2022, 428, 131027. [Google Scholar] [CrossRef]
  40. Cai, L.; Zhao, J.; Li, H.; Park, J.; Cho, I.S.; Han, H.S.; Zheng, X. One-step hydrothermal deposition of Ni:FeOOH onto photoanodes for enhanced water oxidation. ACS Energy Lett. 2016, 1, 624–632. [Google Scholar] [CrossRef]
  41. Wang, X.; Liu, W.; Wang, J.; Li, C.; Zheng, R.; Zhang, H.; Liu, J.; Zhang, X. Cobalt and vanadium co-doped FeOOH nanoribbons: An iron-rich electrocatalyst for efficient water oxidation. Mater. Chem. Front. 2021, 5, 6485–6490. [Google Scholar] [CrossRef]
  42. She, H.; Yue, P.; Huang, J.; .Wang, L.; Wang, Q. One-step hydrothermal deposition of F:FeOOH onto BiVO4 photoanode for enhanced water oxidation. Chem. Eng. J. 2020, 392, 123703. [Google Scholar] [CrossRef]
  43. Zhang, R.; Ning, X.; Wang, Z.; Zhao, H.; He, Y.; Han, Z.; Du, P.; Lu, X. Significantly promoting the photogenerated charge separation by introducing an oxygen vacancy regulation strategy on the FeNiOOH co-catalyst. Small 2022, 18, 2107938. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Zhang, J.; Balogun, M.S.; Tong, Y.; Huang, Y. Oxygen vacancy–based metal oxides photoanodes in photoelectrochemical water splitting. Mater. Today Sustain. 2022, 18, 100118. [Google Scholar] [CrossRef]
  45. Park, G.; Kim, Y.I.; Kim, Y.H.; Park, M.; Jang, K.Y.; Song, H.; Nam, K.M. Preparation and phase transition of FeOOH nanorods: Strain effects on catalytic water oxidation. Nanoscale 2017, 9, 4751−4758. [Google Scholar] [CrossRef]
  46. Mukai, K.; Suzuki, T.M.; Uyama, T.; Nonaka, T.; Morikawa, T.; Yamada, I. High-pressure synthesis of ε-FeOOH from β-FeOOH and its application to the water oxidation catalyst. RSC Adv. 2020, 10, 44756−44767. [Google Scholar] [CrossRef] [PubMed]
  47. Feng, H.; Tang, J.; Tang, W.; Tang, L.; Yu, J.; Wang, J.; Yang, Y.; Ni, T.; Li, C.; Luo, J. A strategy to improve electrochemical water oxidation of FeOOH by modulating the electronic structure. Appl. Mater. Today 2021, 25, 101252. [Google Scholar] [CrossRef]
  48. Zhang, B.B.; Yu, S.Q.; Dai, Y.; Huang, X.J.; Chou, L.J.; Lu, G.X.; Dong, G.J.; Bi, Y.P. Nitrogen-incorporation activates NiFeOx catalysts for efficiently boosting oxygen evolution activity and stability of BiVO4 photoanodes. Nat. Commun. 2021, 12, 6969. [Google Scholar] [CrossRef] [PubMed]
  49. Suryawanshi, M.P.; Ghorpade, U.V.; Shin, S.W.; Suryawanshi, U.P.; Shim, H.J.; Kang, S.H.; Kim, J.H. Facile, Room Temperature, Electroless Deposited (Fe1−x,Mnx)OOH Nanosheets as Advanced Catalysts: The Role of Mn Incorporation. Small 2018, 14, 1801226. [Google Scholar] [CrossRef]
  50. Wang, J.C.; Zhang, Y.; Bai, J.; Li, J.H.; Zhou, C.H.; Li, L.; Xie, C.Y.; Zhou, T.S.; Zhu, H.; Zhou, B.X. Ni doped amorphous FeOOH layer as ultrafast hole transfer channel for enhanced PEC performance of BiVO4. J. Colloid Interface Sci. 2023, 644, 509–518. [Google Scholar] [CrossRef]
  51. He, Y.R.; Zhang, R.F.; Wang, Z.; Ye, H.Q.; Zhao, H.H.; Lu, B.Z.; Du, P.Y.; Lu, X.Q. Unveiling the Influence of Sulfur Doping on Photoelectrochemical Performance in BiVO4/FeOOH Heterostructures. Anal. Chem. 2024, 96, 100–116. [Google Scholar] [CrossRef]
  52. Shi, J.; Hara, Y.; Sun, C.; Anderson, M.A.; Wang, X. Three-Dimensional High-Density Hierarchical Nanowire Architecture for High-Performance Photoelectrochemical Electrodes. Nano Lett. 2011, 11, 3413–3419. [Google Scholar] [CrossRef]
  53. Wang, H.F.; Yuan, M.L.; Zhang, J.X.; Bai, Y.L.; Zhang, K.; Li, B.; Zhang, G.J. Rational element-doping of FeOOH-based electrocatalysts for efficient ammonia electrosynthesis. EES Catal. 2024, 2, 324–334. [Google Scholar] [CrossRef]
  54. Wang, R.Q.; Li, D.Y.; Wang, H.L.; Liu, C.L.; Xu, L.J. Preparation, Characterization, and Performance Analysis of S-Doped Bi2MoO6 Nanosheets. Nanomaterials 2019, 9, 1341. [Google Scholar] [CrossRef]
  55. 55 Faria, D.L.A.D.; Silva, S.V.; Oliveira, M.T.D. Raman Microspectroscopy of Some Iron Oxides and Oxyhydroxides. J. Raman Spectrosc. 1997, 28, 873–878. [Google Scholar] [CrossRef]
  56. Wang, H.L.; Zhao, Z.F.; Xu, Z.K.; Li, L.; Lin, S.Y. Efficient and durable S-doped Ni/FeOOH electrocatalysts for oxygen evolution reactions. Dalton Trans. 2023, 52, 1113–1121. [Google Scholar] [CrossRef]
  57. Quang, N.D.; Van, P.C.; Majumder, S.; Jeong, J.R.; Kim, D.; Kim, C. Rational construction of S-doped FeOOH onto Fe2O3 nanorods for enhanced water oxidation. J. Colloid Interface Sci. 2022, 616, 749–758. [Google Scholar] [CrossRef] [PubMed]
  58. Luo, Z.B.; Li, C.C.; Liu, S.S.; Wang, T.; Gong, J.L. Gradient doping of phosphorus in Fe2O3 nanoarray photoanodes for enhanced charge separation. Chem. Sci 2017, 8, 91–100. [Google Scholar] [CrossRef]
  59. Zhu, W.W.; Wei, Y.Q.; Liu, Z.C.; Zhang, Y.C.; He, H.C.; Yang, S.G.; Li, Z.D.; Zou, Z.G.; Zhou, Y. Construction of unique heterojunction photoanodes through in situ quasi-epitaxial growth of FeVO4 on Fe2O3 nanorod arrays for enhanced photoelectrochemical performance. Catal. Sci. Technol. 2022, 12, 4372–4379. [Google Scholar] [CrossRef]
  60. Zhu, Y.; Zhao, X.; Li, J.; Zhang, H.; Chen, S.; Han, W.; Yang, D. Surface modification of hematite photoanode by NiFe layered double hydroxide for boosting photoelectrocatalytic water oxidation. J. Alloy Compd. 2018, 764, 341–346. [Google Scholar] [CrossRef]
  61. Wu, Q.; Meng, D.; Zhang, Y.; Zhao, Q.; Bu, Q.; Wang, D.; Zou, X.; Lin, Y.; Li, S.; Xie, T. Acid-treated Ti4+ doped hematite photoanode for efficient solar water oxidation—Insight into surface states and charge separation. J. Alloys Compd. 2019, 782, 943–951. [Google Scholar] [CrossRef]
Scheme 1. Sketch map of the synthesis procedure for the photoanodes.
Scheme 1. Sketch map of the synthesis procedure for the photoanodes.
Nanomaterials 15 00767 sch001
Figure 1. SEM images of the (a) Fe2O3/FTO, (b) FeOOH/Fe2O3/FTO, and (c) S:FeOOH/Fe2O3/FTO photoanodes. (d) XRD patterns of the prepared photoanodes.
Figure 1. SEM images of the (a) Fe2O3/FTO, (b) FeOOH/Fe2O3/FTO, and (c) S:FeOOH/Fe2O3/FTO photoanodes. (d) XRD patterns of the prepared photoanodes.
Nanomaterials 15 00767 g001
Figure 2. TEM (a) and HRTEM (b) images of S:FeOOH/Fe2O3.
Figure 2. TEM (a) and HRTEM (b) images of S:FeOOH/Fe2O3.
Nanomaterials 15 00767 g002
Figure 3. XPS spectra: (a) Fe 2p, (b) O 1s, (c) S 2p.
Figure 3. XPS spectra: (a) Fe 2p, (b) O 1s, (c) S 2p.
Nanomaterials 15 00767 g003
Figure 4. (a) LSV curves of S:FeOOH/Fe2O3 photoanodes with different reaction times for S doping; (b) LSV curves, (c) Tafel plots, and (d) Butler plots of different photoanodes.
Figure 4. (a) LSV curves of S:FeOOH/Fe2O3 photoanodes with different reaction times for S doping; (b) LSV curves, (c) Tafel plots, and (d) Butler plots of different photoanodes.
Nanomaterials 15 00767 g004
Figure 5. (a) J-t plots, (b) ABPEs, and (c) IPCEs of different photoanodes at 1.23 V vs. RHE. (d) Stability tests at 1.23 V under continuous illumination.
Figure 5. (a) J-t plots, (b) ABPEs, and (c) IPCEs of different photoanodes at 1.23 V vs. RHE. (d) Stability tests at 1.23 V under continuous illumination.
Nanomaterials 15 00767 g005
Figure 6. (a) LSV curves without and with Na2SO3. (b) ηbulk and (c) ηsurface without Na2SO3 solution.
Figure 6. (a) LSV curves without and with Na2SO3. (b) ηbulk and (c) ηsurface without Na2SO3 solution.
Nanomaterials 15 00767 g006
Figure 7. CV (ac) and Cdl (d) plots of different photoanodes.
Figure 7. CV (ac) and Cdl (d) plots of different photoanodes.
Nanomaterials 15 00767 g007
Figure 8. (a) Nyquist curves, (b) M-S plots, and (c) TPV curves for Fe2O3, FeOOH/Fe2O3 and S:FeOOH/Fe2O3.
Figure 8. (a) Nyquist curves, (b) M-S plots, and (c) TPV curves for Fe2O3, FeOOH/Fe2O3 and S:FeOOH/Fe2O3.
Nanomaterials 15 00767 g008
Figure 9. UPS for the secondary electron cut-off energy and the binding energy region of EVBM of Fe2O3 (a,b) and S:FeOOH (c,d).
Figure 9. UPS for the secondary electron cut-off energy and the binding energy region of EVBM of Fe2O3 (a,b) and S:FeOOH (c,d).
Nanomaterials 15 00767 g009
Figure 10. (a) The band structures of Fe2O3 and S:FeOOH; (b) a sketch map of the charge transfer mechanism for the S:FeOOH/Fe2O3 photoanode.
Figure 10. (a) The band structures of Fe2O3 and S:FeOOH; (b) a sketch map of the charge transfer mechanism for the S:FeOOH/Fe2O3 photoanode.
Nanomaterials 15 00767 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, Y.; Zhang, Y.; Jing, B.; Liu, X.; Wang, D. S-Doped FeOOH Layers as Efficient Hole Transport Channels for the Enhanced Photoelectrochemical Performance of Fe2O3. Nanomaterials 2025, 15, 767. https://doi.org/10.3390/nano15100767

AMA Style

Zhou Y, Zhang Y, Jing B, Liu X, Wang D. S-Doped FeOOH Layers as Efficient Hole Transport Channels for the Enhanced Photoelectrochemical Performance of Fe2O3. Nanomaterials. 2025; 15(10):767. https://doi.org/10.3390/nano15100767

Chicago/Turabian Style

Zhou, Yanhong, Yiran Zhang, Boyang Jing, Xiaoyuan Liu, and Debao Wang. 2025. "S-Doped FeOOH Layers as Efficient Hole Transport Channels for the Enhanced Photoelectrochemical Performance of Fe2O3" Nanomaterials 15, no. 10: 767. https://doi.org/10.3390/nano15100767

APA Style

Zhou, Y., Zhang, Y., Jing, B., Liu, X., & Wang, D. (2025). S-Doped FeOOH Layers as Efficient Hole Transport Channels for the Enhanced Photoelectrochemical Performance of Fe2O3. Nanomaterials, 15(10), 767. https://doi.org/10.3390/nano15100767

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop